• Aucun résultat trouvé

Quasi-explicit time-integration schemes for dynamic fracture with set-valued cohesive zone models

N/A
N/A
Protected

Academic year: 2021

Partager "Quasi-explicit time-integration schemes for dynamic fracture with set-valued cohesive zone models"

Copied!
29
0
0

Texte intégral

(1)

HAL Id: hal-00736779

https://hal.archives-ouvertes.fr/hal-00736779

Submitted on 30 Sep 2012

HAL is a multi-disciplinary open access archive for the deposit and dissemination of sci- entific research documents, whether they are pub- lished or not. The documents may come from teaching and research institutions in France or abroad, or from public or private research centers.

L’archive ouverte pluridisciplinaire HAL, est destinée au dépôt et à la diffusion de documents scientifiques de niveau recherche, publiés ou non, émanant des établissements d’enseignement et de recherche français ou étrangers, des laboratoires publics ou privés.

Quasi-explicit time-integration schemes for dynamic fracture with set-valued cohesive zone models

David Doyen, Alexandre Ern, Serge Piperno

To cite this version:

David Doyen, Alexandre Ern, Serge Piperno. Quasi-explicit time-integration schemes for dynamic fracture with set-valued cohesive zone models. Computational Mechanics, Springer Verlag, 2013, 52 (2), pp.401-416. �10.1007/s00466-012-0819-2�. �hal-00736779�

(2)

(will be inserted by the editor)

Quasi-explicit time-integration schemes for dynamic fracture with set-valued cohesive zone models

D. Doyen · A. Ern · S. Piperno

Received: date / Accepted: date

Abstract We investigate quasi-explicit time-integration schemes for solving dy- namic fracture problems with set-valued cohesive zone models. These schemes combine a central difference time-integration scheme and a partially implicit and lumped treatment of the cohesive forces. At each time step, the displacements of the nodes in the interior of the domain are computed in an explicit way, while the displacements of each node at the interface are computed by solving a local nonlinear problem. The method provides a general and robust way of treating the set-valued cohesive zone model while keeping a moderate computational cost.

Keywords cohesive zone model·finite elements·time-integration scheme

1 Introduction

Cohesive zone models have been introduced in the late 50s [1, 2, 11]. They can be applied to a large range of materials (concrete, steel, etc...) and fracture processes (brittle fracture, ductile fracture, fatigue, dynamic fracture) and they can be easily enriched with more complex physical behaviors (contact and friction after deco- hesion, corrosion, etc...). Moreover, cohesive zone models fit quite well within the

This work was partly supported by Electricit´e de France (EdF R&D) through a CIFRE PhD fellowship.

D. Doyen

Universit´e Paris-Est, Laboratoire d’Analyse et de Math´ematiques Appliqu´ees (UMR 8050), UPEMLV, UPEC, CNRS, F-77454, Marne-la-Vall´ee, France

E-mail: david.doyen@univ-mlv.fr A. Ern

Universit´e Paris-Est, CERMICS, Ecole des Ponts ParisTech, 77455 Marne la Vall´ee Cedex 2, France

E-mail: ern@cermics.enpc.fr S. Piperno

Universit´e Paris-Est, CERMICS, Ecole des Ponts ParisTech, 77455 Marne la Vall´ee Cedex 2, France

E-mail: serge.piperno@enpc.fr

(3)

framework of finite elements. For all these reasons, they are now widely used in engineering simulations. A cohesive zone model can describe the mechanical forces along a fracture — it can be simply viewed as a boundary condition. The interface forces depend at least on the opening (displacement jump at the interface). In a typical cohesive zone model, the separation occurs at the interface only after a critical stress has been reached. When the separation has occurred, cohesive forces remain. These forces decrease when the opening increases and tend to van- ish (softening behavior). Physically, the cohesive forces represent the weakening of the material in the fracture process zone ahead of the crack tip. Furthermore, real cracks cannot experience self-healing in general. To take into account this ir- reversibility, one can introduce a history parameter, such as the maximal opening.

For quasi-static fracture, cohesive zone models depending on opening and maxi- mal opening are well established and in good agreement with experiments. In the dynamic case, numerical simulations with such cohesive zone models predict of- ten crack speeds far higher than those observed in the experiments. For instance, for mode-I fracture in brittle materials, numerical crack tip speeds are close to the Rayleigh wave speed, while experimental crack tip speeds nearly reach half of this value (see for instance [22, 20]). To remedy this, cohesive zone models de- pending on the opening rate have been designed [21, 27]. Such models are called rate-dependent.

Since the crack tip speed is high, typically of the same order as the wave speeds, small time steps are needed to capture accurately the fracture phenomenon.

Therefore, it seems natural to consider an explicit time-integration scheme. For cohesive zone models in which the interface forces are related to the opening by a classical function, the use of an explicit time-integration scheme is straightforward [26]. However, in most cohesive zone models, the interface forces are not related to the opening by a classical function, but by a set-valued map. Indeed, most cohesive zone models feature perfect initial adhesion, contact or rigid unloading.

There are two main difficulties in using fully explicit schemes in such a context.

Firstly, the interface forces are not defined for negative normal opening (see Figure 1, left). Secondly, the interface tangential forces are discontinuous with respect to the tangential opening (see Figure 1, right), and this can cause oscillations. A first option consists in regularizing the set-valued map to turn it into a single- valued map. Unfortunately, the regularization of a non-interpenetration condition deteriorates substantially the stability condition of explicit schemes (the penalty contact condition introduces in the model an artificial stiffness larger than the material stiffness). Moreover, replacing a discontinuity by a very stiff slope does not really solve the problem. Alternatively,ad hocprocedures have also been developed to treat a few specific cases of set-valued interface forces: allowing the separation only after a failure criterion has been reached [5, 19, 12],a posteriorienforcement of the non-interpenetration condition [5, 12], tolerance parameter on the tangential opening [12]. For complex cohesive zone models with several set-valued parts, the combined use of these procedures generally becomes quite intricate, or even unfeasible.

In the present work, we focus on dynamic fracture models where the mate- rial can only crack along a prescribed surface (fracture interface). In other words, the crack path is known in advance. This assumption may appear as a limitation.

However, fracture models predicting the crack path are still quite challenging for in- dustrial applications. Moreover, for a large range of applications (interfacial crack,

(4)

small propagation crack), postulatinga priorithe crack path is reasonable (see [16]

for further discussion). We assume that the bulk behavior is governed by the linear elastodynamic equations and that the separation process at the fracture interface obeys a cohesive zone model. We consider a generic cohesive zone model depending on the opening, the opening rate, and the maximal opening. This generic model encompasses most of the usual cohesive zone models. Space semi-discretization is achieved using P1 finite elements. We propose time-integration schemes that combine a central difference scheme with a partially or fully implicit treatment of the interface forces. The central difference scheme is a standard scheme for elas- todynamics [15]. The implicit treatment of the interface forces provides a general and robust way of treating the set-valued cohesive zone model. In order to keep a moderate computational cost, we use lumping techniques for the mass term and the interface forces. We thus obtain quasi-explicit methods: at each time step, the displacements of the nodes in the interior of the domain are computed in an ex- plicit way, while the displacements of each node at the interface are computed by solving a small nonlinear problem (this can generally be achieved in an analytical way). First, we consider a fully implicit treatment of the interface forces. However, staggering in time the force at the fracture interface can have a sizeable effect on the energy behavior and the accuracy of the time-integration scheme. Conse- quently, we propose a second time-integration scheme, in which the interface forces are split into a set-valued monotone part and a single-valued softening part. The former is treated in an implicit way, the latter in an explicit way. This improves the accuracy and the energy behavior. Note that some of the ad hoc procedures described above can be loosely interpreted as an implicit treatment of the set val- ued part of the cohesive zone model (failure criterion,a posteriori enforcement of the contact condition).

We begin by presenting the generic cohesive zone model and examples which fit into this framework (Section 2). We then formulate the continuous problem of dynamic fracture (Section 3). Sections 4 and 5 are devoted to the finite element discretization in space and to the time-integration schemes, respectively. We dis- cuss the implementation of the schemes in Section 6. Finally, numerical results are presented in Section 7, and conclusions are drawn in Section 8.

2 Cohesive zone model

We consider a generic cohesive zone model. The forces at the fracture interface are described by a set-valued map which depends on the opening, the opening rate, and the maximal effective opening (the notion of effective opening is defined below).

2.1 Generic model

Let (·,·) denote the usual scalar product in Rn (n 1) and let | · | denote the corresponding Euclidean norm. LetP(Rn) denote the set of all subsets ofRn. In a d-dimensional problem (d= 2 ord= 3), the cohesive zone model is characterized by a set-valued mapR:Rd×Rd×Rd→ P(Rd). The arguments ofRare the max- imal effective opening, the opening rate, and the opening, respectively. For each

(5)

triplet (δ, z, p), the mapR(δ, z, p) yields a set of vectors, which are the admissible interaction forces. The first component ofλR(δ, z, p) is the normal force at the interface and the second and third ones are the tangential forces. For an opening p R3 (resp. p R2), the effective opening is defined as ¯δ(p) = (p1,|p2|,|p3|) (resp. ¯δ(p) = (p1,|p2|)).

Since a cohesive zone model describes a softening behavior, the set-valued map Ris not monotone with respect to the opening. However, the slope of the softening part ofR is assumed to be bounded. This assumption is, in particular, useful to establish the well-posedness of our first time-integration scheme (see Proposition 1).

Assumption 1 The operatorRsatisfies the following one-sided Lipschitz condition:

there is a real numberCLsuch that, for allδRd, for allzRd, for allp,qRd, pλq, pq)≥ −CL|pq|2, ∀λpR(δ, z, p), ∀λqR(δ, z, q). (1) The present generic model encompasses for instance the Camacho-Ortiz law [5] and the Talon-Curnier law [23], but not the rectangular law (because of the infinite slope of the softening part).

Remark 1 In most cohesive zone models, the operatorR(δ, z,·) is built as the dif- ferential (in a generalized sense) of an energy. This operator being non-monotone, the associated cohesive energy is non-convex.

Remark 2 After decohesion, contact closure can lead to friction phenomena at the interface, which can play an important role in the fracture process. Cohesive zone models including friction have been proposed for instance in [18]. The friction force is generally a monotone set-valued function of the tangential velocity at the interface and could be easily added to our generic model.

2.2 Examples

This section collects some examples of cohesive zone models fitting the above framework. The first two examples can be viewed as simplified variants of the Camacho-Ortiz law [5].

A reversible triangular model with uncoupled normal and tangential forces This model depends on the opening and prescribes uncoupled normal and tangential interface forces. It relies on two parameters:σc, the maximal cohesive force, anddc, the critical opening. It can be represented by a set-valued map R:Rd → P(Rd) whose components are independent. The normal component R1 : R → P(R) is such that

R1(p) :=

(−∞, σc] ifp= 0, σc

1dp

c

if 0< pdc, 0 ifdc< p,

ifp <0.

(6)

dc

σc

λ1

p1

λ2

p2

σc

dc

σc

Fig. 1 Triangular model with uncoupled normal and tangential forces. Normal force (left).

Tangential force (right).

For simplicity, in the definition of cohesive zone models, a singleton{x}is simply denoted x. The tangential components R2 :R → P(R) and R3 : R→ P(R) are such that

R2(p) =R3(p) :=

0 ifp <−dc,

−σc

1 +dp

c

if dcp <0, [−σc, σc] ifp= 0,

σc

1 dp

c

if 0< pdc, 0 ifdc< p.

This model is represented in Figure 1. It is easy to check that this model satisfies Assumption 1 with CL = σc/dc. Moreover, energies Ψ1 : R+ R, Ψ2 :R R, andΨ3:RRcan be associated with this model, namely

Ψ1(p) :=

(σcp 12dp

c

if 0pdc,

1

2σcdc ifdc< p, and

Ψ2(p) =Ψ3(p) :=

1

2σcdc ifp <−dc,

−σcp

1 +2dp

c

if dcp <0, σcp

12dp

c

if 0pdc,

1

2σcdc ifdc< p.

An irreversible triangular model with only normal force This model depends on the normal openingpand maximal effective normal openingδ and prescribes only the normal force. Moreover, it is irreversible with a linear unloading. As the previous model, it involves two parameters: σc, the maximal cohesive force, and dc, the critical opening. It can be represented by the set-valued mapRirr1 :R×R→ P(R)

(7)

dc

σc

λ

δ p dc

σc

λ

δ p

Fig. 2 An irreversible triangular model with only normal force. Linear unloading (left). Rigid unloading (right).

such that

Rirr1 (δ, p) :=

(−∞, σc] ifδ=p= 0, (−∞,0] if 0 =p < δ, σc

1dδ

c

p

δ if 0< pδdc, σc

1dp

c

if 0< δ < pdc, 0 ifdc< p, 0δ, 0 ifdc< δ,0p,

otherwise.

This model is represented in Figure 2, left. It satisfies Assumption 1 withCL = σc/dc. A common variant of this model consists in replacing the linear unloading with a rigid unloading (Figure 2, right). An energy Ψ1irr : R+×R+ R and a dissipated energy ˜Ψ1irr : R+ R can be associated with the irreversible model with linear unloading. They are defined as follows:

Ψ1irr(δ, p) :=

σc

δ2+p 2dp2

c

if 0< δ < pdc, σc

1dδ

c

p2

if 0< pδdc,

1

2σcdc ifdc< p, δ= 0, 0 ifdc< δ,0p, Ψ˜1irr(δ) :=

(1

2σcδ if 0δdc,

1

2σcdc ifdc< δ.

A rate-dependent triangular model with only normal force [27] This model depends on the normal openingpand the normal opening ratez, and prescribes only the normal interface force. It relies on three parameters: σc, the maximal cohesive force,dc, the critical opening, andη, a viscosity parameter. It can be represented by the set-valued mapR1visc:R×R→ P(R) such that

Rvisc1 (z, p) :=

(−∞, σc] ifp= 0, σc

1d p

c(1+ηz+)

if 0< p < dc(1 +ηz+),

0 ifdc(1 +ηz+)p,

otherwise,

(8)

λ

dc d˜c p σc

Fig. 3 A rate-dependent triangular model with only normal force ( ˜dc:=dc(1 +ηz+)).

wherez+ denotes the positive part ofz. This model is represented in Figure 3. It satisfies Assumption 1 withCL =σc/dc.

2.3 Link with Griffith’s model

It is possible to make a link between cohesive zone models and Griffith’s models.

When cohesive forces act over a sufficiently short range, the stress fields near the crack tip are equivalent in both models. Furthermore, the material parameter used in Griffith’s model, the fracture toughness Gc, is equal to the energy needed to completely open the crack in the cohesive zone model. A formal argument for this asymptotic analysis can be found in [25] and a rigorous proof for a simple model in [17]. The fracture toughness corresponding to the rate-independent triangular models (presented above) is

Gc= 1 2σcdc.

In the rate-dependent triangular model, the fracture toughness increases with the opening rate.

3 Continuous problem

We now formulate the governing equations of the dynamic fracture problem.

3.1 Geometry

We consider a domain Rd (d= 2 or d= 3) and we assume that the crack can only appear on a (d1)-dimensional smooth surface Γ (see Figure 4). We call Γ the fracture interface. We set ˜ := \Γ. We can fix an orientation and define two sides for Γ, a positive side and a negative side. Let v : Rd be a displacement field. The trace of v on the positive side is denotedv+, the trace on the negative side is denoted v. We denote ν the unit normal vector to Γ pointing to the positive side. We define two tangential unit vectors τ1 andτ2, so that (ν, τ1, τ2) forms a local direct orthonormal basis (obviously, ford= 2, only

(9)

ΓD ΓN

Ω˜ − +

Γ ν

Fig. 4 Geometric setup.

one tangential vector is considered). The displacement jump ofv at the interface is defined as

[[v]] =v+v. (2)

To define the interface forces, we take into account the local orientation of the interface by introducing the rotation matrixQtransforming the canonical basis of Rd into (ν, τ1, τ2).

3.2 Governing equations

The material is supposed to be linear isotropic elastic with Young modulus E, Poisson ratio νP, and mass density ρ. The elasticity tensor is denoted A. An external loadf is applied to the body. Letu:×(0, T)Rd,(u) :×(0, T) Rd,d, andσ(u) :×(0, T)Rd,dbe the displacement field, the linearized strain tensor, and the stress tensor, respectively. Denoting time-derivatives by dots, the momentum conservation equation reads

ρ¨udivσ=f, σ=A:, = 1

2(∇u+T∇u) in ˜×(0, T). (3) The boundary∂Ωofis partitioned into two disjoint subsetsΓDandΓN. Dirich- let and Neumann conditions are prescribed onΓD andΓN, respectively,

u=uD on ΓD×(0, T), σ·ν=fN onΓN×(0, T). (4) OnΓ, the cohesive law is enforced

σ(u)·ν=−σ(u+)·ν=:λ, λTQR(δ, Q[[ ˙u]], Q[[u]]). (5) The maximal effective openingδ is defined, fort >0, by

δ(t) = sup

s∈[0,t)

¯δ(Q[[u(s)]]). (6)

At the initial time, the displacement, the velocity, and the maximal effective open- ing are prescribed:

u(0) =u0, u(0) =˙ v0, δ(0) =δ0. (7) Equations (3)-(5) can be written in a variational form: seek u such that, for all test functionv,

Z

˜

ρ¨u·v+ Z

˜

(u) :A:(v) = Z

˜

f(t)·v+ Z

ΓN

fN(t)·v Z

Γ

λ·[[v]], (8) whereλTQR(δ, Q[[ ˙u]], Q[[u]]).

(10)

3.3 Mathematical aspects

The mathematical analysis of Problem (3)-(7) is beyond the scope of the present work. However, let us mention some related results.

In the quasi-static case with a reversible cohesive zone model (with perfect adhesion or not), the existence is proven. The solution is in general not unique [10].

In the quasi-static case with an irreversible cohesive zone model, the existence of a solution is proven in [8, 4].

In the dynamic case, it should be possible to prove, using compactness ar- guments similarly to [9], existence for a visco-elastic material and a reversible cohesive law with perfect adhesion, and even to prove existence and uniqueness for an elastic material and a regularized cohesive law.

3.4 Length and time scales

In order to capture accurately a phenomenon with numerical simulations, it is important to choose a time step and a mesh size which resolve its time and length scales. In a dynamic fracture problem, the relevant length scale is the lengthlcoh

of the cohesive zone (the part of the interface which is not completely cracked and where cohesive forces act). The relevant time scale is the crack tip speed divided by the cohesive zone length. Several methodologies have been proposed in the literature to estimate the cohesive zone length; see [24] and references therein.

They all yield, in the case of plane strain and triangular cohesive zone models, an estimation of the form

lcoh=M E 1νP2

Gc

σc2, (9)

where M is a parameter close to 1. For an isotropic linear elastic material and a Griffith model of fracture, a theoretical analysis predicts that the limiting crack tip speed for a mode-I fracture is the Rayleigh wave speed [13, 3]. For mode-II and mode-III fractures, the limiting speeds are the dilatational wave speed and the shear wave speed, respectively [13, 3]. The dilatational and shear wave speeds are given by the following formulae:

cd= s

E(1νP)

ρ(1 +νP)(1P), cs= s

E

2ρ(1 +νP). (10) The Rayleigh wave speed can be estimated by the following expression [13]

cRcs0.862 + 1.14νP

1 +νP

. (11)

Rate-dependent cohesive zone models involve an additional time scale, linked to the opening rate and generally smaller than the time scale linked to the crack tip speed. At least in the quasi-static evolution, analytical estimations of the opening time are provided in [7].

(11)

4 Finite element discretization

In this section, we describe the space approximation of the dynamic fracture prob- lem. Linear finite elements are used together with lumping of the mass term and the interface forces.

4.1 Finite element spaces

In 2D (resp. in 3D), the domain is approximated by a polygon (resp. a poly- hedron) h and the interfaceΓ by a polygonal curve (resp. a polygon)Γh. The domainh is meshed with triangles (resp. tetrahedra) conforming to the interface Γh. LetThdenote the mesh overhand letFhcollect the faces located onΓh. Let ˜h=h\Γh. Let{xi}i∈N be the nodes of the meshThwhereN collects the node indices. Let ND be the indices of nodes where a Dirichlet condition is enforced and byNcthe indices of nodes lying onΓh. The displacements are approximated withP1finite elements:

Vh={vhC0( ˜h)d; vh|T (P1)d, ∀T ∈ Th, andvh(xi) = 0, ∀i∈ ND}.

Note that functions in Vh can be discontinuous across Γh. We consider the La- grange nodes ofVh and denote themi}i∈NLwhereNL collects the correspond- ing node indices. The Lagrange nodes are not exactly the mesh nodes {xi}i∈N

because of the discontinuity at the interface. Specifically, for each mesh node xi

lying on Γh, there are two Lagrange nodes ξi+ and ξi. For all i ∈ Nc, for all vh∈ Vh, we set

vhi+) :=v+h(xi), vhi) :=vh(xi). (12) The cohesive forces are also approximated byP1finite elements,

Lh=hC0h)d; λh|F (P1)d, ∀F ∈ Fh}.

At each nodexilying onΓh, we define normal and tangential unit vectors (νi, τ1i, τ2i) forming a direct orthonormal basis. LetQi be the associated rotation matrix. We define also the set-valued operatorRi:

Ri(·,·,·) =TQiR(·, Qi·, Qi·). (13)

4.2 Lumping of the mass term and the cohesive term

The mass term and the cohesive forces term are lumped. Mass lumping is usual with explicit time-integration schemes. It yields an easy-to-invert mass matrix at each time step, while improving the CFL condition [15]. A way of lumping the mass term is to evaluate it with an approximate quadrature whose Gauss points are the nodes of the finite element space. ForP1 finite elements, it is usual to use the following quadrature formula (in dimensiond)

Z

T

f

d+1

X

i=1

|T|

d+ 1fi), (14)

(12)

where |T| is the measure of the simplex T and i}1≤i≤d+1 its vertices. This quadrature is second-order accurate. The lumped mass term ˆmh:Vh× VhRis built with this quadrature by setting

ˆ

mh(vh, wh) = X

i∈NL

µivhi)·whi), (15) where

µi= X

T∈Ti

ρ|T|/(d+ 1), (16)

Ti being the set of elements for whichξi is a vertex. The lumped cohesive term ˆbh:Lh× VhRis such that

ˆbhh, wh) = X

i∈Nc

βiλh(xi)·

vh(xi)v+h(xi)

, (17)

where

βi= X

F∈Fi

|F|/d, (18)

Fi being the set of faces for which xi is a vertex and |F| the measure of the faceF. Finally, the stiffness term ah :Vh× Vh R and the external force term lh: [0, T]× VhRare built in a standard way, namely

ah(vh, wh) :=

Z

˜h

(vh) :A:(wh), (19)

lh(t, vh) :=

Z

˜h

f(t)·vh+ Z

ΓN

fN(t)·vh, (20) up to quadratures forlh. We define the matrices ˆMh,Kh, and ˆBhassociated with the bilinear forms ˆmh, ah, and ˆbh, respectively. We also define Lh(t) to be the column vector associated with the linear formlh(t,·). Foruh ∈ Vh, we define Uh

as the column vector whose components are the coordinates of uh in the finite element basis. We denote NV the size of Uh. We denote Uh,i the d-dimensional sub-vector associated with the Lagrange nodeξi. Similarly, forλh∈ Lh, we define Λh as the column vector whose components are the coordinates ofλh in the finite element basis. We denote NΛ the size of Λh. We denote Λh,i the d-dimensional sub-vector associated with the nodexi. Finally, we define, for alli∈ Nc,

[[Uh]]i =Uh,i+Uh,i and {Uh}i= Uh,i++Uh,i

2 . (21)

For eachi∈ NL, we denote respectively ˆMh,i andKh,i thed×dsub-matrices of Mˆh andKh associated with the Lagrange node ξi. For eachi ∈ Nc, we denote Bˆh,i thed×dsub-matrix of ˆBh associated with the nodexi. We define the set- valued operatorRh :RNΛ×RNΛ×RNΛ → P(RNΛ) such that for allh RNΛ, ZhRNΛ,PhRNΛ,

ΛhRh(∆h, Zh, Ph)⇐⇒Λh,iRi(∆h,i, Zh,i, Ph,i)∀i∈ Nc. (22)

(13)

The space semi-discrete problem takes the form

MˆhU¨h(t) +KhUh(t) =Lh(t) + ˆBhΛh(t), (23) Λh(t)Rh(∆h(t),[[ ˙Uh(t)]],[[Uh(t)]]), (24)

h,i(t) = sup

s∈[0,t)

δ(Q¯ i[[Uh(s)]]i), ∀i∈ Nc. (25) Remark 3 In the present work, we consider P1 finite elements. Other types of finite elements can be used, provided an accurate lumping technique is available.

For instance, this is the case forQ1 elements (see [15]). ForPk andQk elements with k 2, the lumping techniques are more subtle (see for instance [6] and references therein).

Remark 4 Because of lumping, even when the nodes of the interface are in the perfect adhesion regime, our discretization is not equivalent to a discretization without interface. Consequently, small spurious wave reflections can occur at the interface in the perfect adhesion regime.

5 Time-integration schemes

It remains now to discretize the problem in time. The time-integration schemes we propose are based on the central difference scheme. To begin with, let us recall the main properties of this scheme in the linear elastodynamic case. We then describe and analyze two schemes for the dynamic fracture problem.

5.1 Central differences for elastodynamics

For simplicity, the interval [0, T] is divided into equal subintervals of length ∆t.

We set tn = n∆t and denote Uhn the approximation of Uh at time tn. For the central difference scheme, the discrete velocity and the discrete acceleration are defined respectively as

U˙hn:= Uhn+1Uhn−1

2∆t and U¨hn:= Uhn+12Uhn+Uhn−1

∆t2 . (26)

At each time step of the central difference scheme, one seeksUhn+1 such that 1

∆t2Mˆh(Uhn+12Uhn+Uhn−1) +KhUhn=Lh(tn). (27) The central difference scheme exhibits a stability condition (CFL condition) of the form

cd∆tO(hmin), (28)

where hmin is the smallest mesh element diameter. An admissible value of the constant in the CFL condition can be specified in 1D and for structured meshes in higher dimension. The elastic energy, the kinetic energy, and the total energy of the discrete system at timetnare respectively defined by

Eeln := 1

2(KhUhn, Uhn), Ekinn := 1 2

MˆhU˙hn,U˙hn

, En:=Eeln+Enkin. (29)

(14)

In linear elastodynamics with no forcing termLh(tn), the central difference scheme does not preserve the energy. Nevertheless, the scheme preserves the following quadratic form, referred to as a shifted energy,

En:=En∆t2 8

MˆhU¨hn,U¨hn

. (30)

With a forcing term, the following shifted energy balance holds true:

En+1− En= 1

2(Lh(tn+1) +Lh(tn), Uhn+1Uhn). (31)

5.2 Scheme A (fully implicit interface forces)

The scheme A combines a central difference scheme with an implicit treatment of the interface forces. More precisely, the interface forces are implicit in the opening, while they are explicit in the opening rate and in the maximal effective opening.

Scheme A.SeekUhn+1RNV andΛn+1h RNΛ such that

1

∆t2

Mˆh(Uhn+12Uhn+Uhn−1) +KhUhn=Lh(tn) + ˆBhΛn+1h , (32) Λn+1h Rh(∆nh, Zhn,[[Uhn+1]]), (33) where, for alli∈ Nc,

Zh,in = [[Uhn]]i[[Uhn−1]]i

∆t , nh,i= max(∆n−1h,i ,δ(Q¯ i[[Uhn]]i)). (34) A way of implementing this scheme will be described in Section 6.1. We now prove that, at each time step, the problem is well-posed under a mild restriction on the time step. We observe that condition (36) below is indeed mild, since for hcandhmin of the same size and small enough, the stability condition (28) of the central difference scheme imposes a more stringent limit on the time step than the well-posedness condition (36).

Proposition 1 (Well-posedness)Problem (32)-(34) has a unique solution un- der the conditions

µi+

∆t2 >2CLβi and µi

∆t2 >2CLβi, ∀i∈ Nc, (35) where the coefficientsµi andβi are defined by (16) and (18). For a quasi-uniform mesh, the above condition can be rewritten as

∆t2 hc

< C, (36)

where hc is the mesh size at the interface and C a constant independent of the mesh size.

Références

Documents relatifs

Properties of cast-nephropathy light chains were more heteroge- neous but clearly differed from those of Fanconi's syndrome: (i) 9 out of 12 were of the A-type; (ii) only four yielded

Using photoemission and transport measurements, we have confirmed the degenerately n-type doping of the material and are able to propose an electronic spectrum in absolute

Pulsed experiments typically are at full load a very small percentage of the lifetime of the experimental program. Even though FIRE and IGNITOR are resistive machines, they are

• In case of coupled effect between the rapid varying moisture on the FPZ and the viscoelastic mechano-sorptive behavior associated to the shrinkage-swelling on the crack growth

A Cranz-Schardin camera was used to record images on film of the crack tip region before and during the dynamic impact of a standard three-point bend fracture specimen

To sum up, the experimental device is designed to pro- duce a fracture in a thin layer (thickness, h) of granular material (glass beads of typical radius r) in contact with a

This study described the aerosol optical thicknesses encoun- tered during the great European heat wave of August 2003 by comparing data derived from the POLDER-2 satellite sensor,

L’accès à ce site Web et l’utilisation de son contenu sont assujettis aux conditions présentées dans le site LISEZ CES CONDITIONS ATTENTIVEMENT AVANT D’UTILISER CE SITE